Predictions of Binding for Dopamine D2 Receptor ... - ACS Publications

National Center for High-performance Computing, Hsin-Shi, Tainan County 741, ... Kaohsiung 804, Taiwan, and The Department of Information Management, ...
0 downloads 4 Views 3MB Size
J. Chem. Inf. Model. 2009, 49, 2369–2375

2369

Predictions of Binding for Dopamine D2 Receptor Antagonists by the SIE Method Yeng-Tseng Wang,*,†,‡ Zhi-Yuan Su,§ Chang-Huain Hsieh,† and Cheng-Lung Chen*,‡ National Center for High-performance Computing, Hsin-Shi, Tainan County 741, Taiwan, and The Department of Chemistry, National Sun Yat-Sen University, Kaohsiung 804, Taiwan, and The Department of Information Management, Chia Nan University of Pharmacy & Science, Tainan 717, Taiwan Received June 25, 2009

The control of tetralindiol derivative antagonists released through the inhibition of dopamine D2 receptors has been identified as a potential target for the treatment of schizophrenia. We employed molecular dynamics simulation techniques to identify the predicted D2 receptor structure. Homology models of the protein were developed on the basis of crystal structures of four receptor crystals. Compound docking revealed the possible binding mode. In addition, the docking analyses results indicate that five residues (Asp72, Val73, Cys76, Leu183, and Phe187) were responsible for the selectivity of the tetralindiol derivatives. Our molecular dynamics simulations were applied in combination with the solvated interaction energies (SIE) technique to predict the compounds’ docking modes in the binding pocket of the D2 receptor. The simulations revealed satisfactory correlations between the calculated and experimental binding affinities of all seven tetralindiol derivative antagonists, as indicated by the obtained R2 value of 0.815. INTRODUCTION

The dopamine molecule has been associated with many physiological functions such as fine movement coordination, cognition, emotion, and memory by the pituitary reward system.1 Alterations in dopaminergic function are involved in the pathogenesis of Parkinson’s disease,2 psychomotor diseases, and schizophrenia.3 Dopamine receptors can be classified into two major subfamilies, D1 and D2 receptors, according to their pharmacological and domain functional characteristics.4 Therefore, the dopamine receptor families have been targeted in the development of treatment medications for these disorders. The dopamine D2 receptor is a G protein-coupled receptor located on postsynaptic dopaminergic neurons that is centrally involved in reward-mediating mesocorticolimbic pathways.5 Signaling through D2 receptors governs physiologic functions related to locomotion, hormone production, and drug abuse. D2 receptors are also known targets of antipsychotic drugs used to treat psychomotor diseases such as schizophrenia, a debilitating mental illness that affects 0.5-1.5% of the worldwide population.6 Although the biophysical and pharmacological properties of D2 receptors have been studied,7 many problems remain unresolved due to the lack of three-dimensional structures and experimental limitations. In recent years, a number of antagonists based on the cleavage of native substrates have been found, or designed, to inhibit D2 receptors.1,4,8-18 For instance, stepholidine (SPD) isolated from the Chinese herb Stephania is a potential lead compound against D1 and D2 receptors. The pharmacophore of SPD molecules has been well studied in experi* Corresponding author e-mail: [email protected] (Y.T.W.), [email protected] (C.L.C.). † National Center for High-performance Computing. ‡ National Sun Yat-Sen University. § Chia Nan University of Pharmacy & Science.

ments4 and molecular simulations.19 We chose tetralindiol derivative antagonists for pharmacophore studies of D2 receptors as they possess the strong pharmacokinetic properties of D2 receptors.12 In the present study, tetralindiol derivative8-18 compounds and their experimental biological binding affinities8-18 (IC50) were chosen to simulate D2 receptor pharmacological activities; these compounds are listed in Table 1. The transfer function20 (∆Gbind ) -RT ln(IC50)) is used to transfer the IC50 values to the experimental ∆Gbind values; these experimental ∆Gbind values are listed in Table 5 in the Supporting Information. Considering that the crystal structures of D2 receptors and inhibitors are not available, homology modeling, molecular docking studies, molecular dynamics (MD) simulations, and binding free energy calculations were performed to gain further insight into the binding interactions between D2 receptors and a series of tetralindiol derivative inhibitors. METHODS

Homology Modeling of Dopamine D2 Receptor. The dopamine D2 receptor sequence was collected from the National Center for Biotechnology Information (NCBI) protein database (ID: NP_000786).21 A sequence similarity search for the D2 receptor against other protein databank sequences was performed using the NCBI BLAST server.22 Four receptor structures were identified as homologous from the Protein Data Bank (PDB)23-26 (Table 2). Automated sequence alignment (Figure 1) and template and target analyses were carried out using the Clustal X 2.09 program.27 Meanwhile, the Modeler program package28 was used to model a D2 receptor protein sequence. Four different template sequences were aligned with the D2 receptor sequence, and this alignment was supplied along with the 3D coordinates of the templates as an input to the program. The Modeler implements comparative protein structure

10.1021/ci9002238 CCC: $40.75  2009 American Chemical Society Published on Web 09/21/2009

2370

J. Chem. Inf. Model., Vol. 49, No. 10, 2009

Table 1. Tetralindiol Derivative Antagonists

WANG

Table 2. Templates Used for Dopamine D2 Receptor Homology Modeling protein name human beta2 adrenergic receptor human B2-adrenergic G protein-coupled receptor human B2-adrenergic G protein-coupled receptor human beta2 adrenoceptor receptor

modeling by satisfying spatial restraints.29,30 In order to obtain a relaxed conformation, the three models generated were initially subjected to an energy minimization process by using the conjugate gradient method for about 2000 iterations and to 2-ns isothermal, constant-volume MD simulation, with AMBER FF99 all-hydrogen amino acid parameters in the AMBER31 program running on a LINUX server. To assess the quality of the minimized models, PROCHECK32 analysis was also undertaken. Here, we considered two aspects in choosing the model: the “energy minima” and the optimum binding site to accommodate the D2 receptor inhibitors. Molecular Docking. All tetralindiol derivative compounds were constructed and minimized using the SYBYL33 modeling program. We aligned the model structure and the G protein-coupled receptor.24 We then used the structure of the G protein-coupled receptor inhibitor (CUA: (2S)-1-(9Hcarbazol-4-yloxy)-3-(isopropylamino) propan-2-ol) as a temp-

ET AL.

PDB ID identity resolution reference 3D4S

28%

2.8 Å

11

2RH1

28%

2.4 Å

12

3EML

28%

2.6 Å

13

2R4S

33%

3.4 Å

14

late to generate the receptor active site. The GOLD34 program was used to dock all compounds into the active site of the model structure, which was defined as all residues within 10 Å of the aligned CUA inhibitor. This program, which makes use of hydrogen binding annealing parameters, GOLDscore, and a genetic algorithm, provides a fast and accurate means to dock small compounds into fixed protein binding sites. For every compound, 60 conformations were obtained from docking and then scored by the GOLDscore. The binding modes were identified after sorting the GOLDscore values, and the conformations of the best GOLDscores were selected for subsequent MD simulation. Molecular Dynamics Simulation. Calculations were performed with the AMBER package using the AMBER FF99 all-hydrogen amino acid and general amber force field (GAFF) parameters. The GAFF partial atomic charges are often based on the RESP fitting procedure of the electrostatic potential obtained at the HF/6-31G(d) level of theory.35,36 The B3LYP/6-31G(d,p) level of theory is also accurate and suitable for the electrostatic potential calculations of biomolecules and small organic molecules.37-40 Those levels of theory overestimate the gas-phase partial atomic charges giving rise to effectively polarized force field. The geometries of the seven tetralindiol derivatives and dilauryl-phosphatidyl-ethanol-amine (DLPE) were fully optimized, and their electrostatic potentials were obtained using a single-point calculation; both operations were carried out at the B3LYP level with the 6-31G(d,p) basis set using the GAUSSIAN 03 program.41 Subsequently, their partial charges were obtained by the restrained electrostatic potential (RESP) using Antechamber.42 From the docking simulations, the complex structures were generated and then inserted into the solvents. The solvent molecules contain a DLPE lipid bilayer and water molecules. In addition, the initial compound II complex structures are shown in Figure 2. All MD simulations were performed in the canonical ensemble43 (with a simulation temperature equal to 310 K), unless noted, using a leapfrog integrator, an integration time step of 0.002 ps, and SHAKE44 data for all covalent bonds involving hydrogen atoms. In electrostatic interactions, atom-based truncation was undertaken using the PME method. In addition, the switch van der Waals functions were also used with a 1.8-nm cutoff for atom-pair lists. The complex structures were minimized for 50,000 conjugate gradient steps. The minimized complex structures were then subjected to a 5-ns isothermal, constant volume MD simulation. The final structures from these simulations were used to initiate the functionally important residues and the solvated interaction energies (SIE) calculations. Functionally Important Residues. In a protein, every residue is not equally important. Some residues are essential

DOPAMINE D2 RECEPTOR ANTAGONISTS

J. Chem. Inf. Model., Vol. 49, No. 10, 2009 2371

Figure 1. Dopamine D2 receptor (human) sequence aligned with four different template sequences: 3D4S, human beta2 adrenergic receptor; 2RH1, human B2-adrenergic G protein-coupled receptor; 3EML, human beta2 adrenergic receptor; 2R4S, human beta2 adrenergic receptor.

Figure 2. Overview of compound IIsD2 receptor/DLPE/water system. The D2 receptor is shown in the ribbon; the hydrophobic chains of bilayer lipids are labeled L, and the regions containing water are labeled W. The 3-D size of the whole system is also shown. In the binding site, compound II is amplified and shown in a stick model. Table 3. Data Reorient the GOLDscores by Docking the Seven Tetralindiol Derivatives Compounds into Two Homology Dopamine D2 Receptor Models compound

model 1 (KJ/mol)

model 2 (KJ/mol)

I II III IV V VI VII

NULL NULL NULL NULL NULL NULL NULL

–41.4052 –39.3230 –42.6992 –41.6285 –37.3118 –41.4983 –44.8341

for important functions and structure stability and are known as conserved residues. However, others can be replaced by means of evolution. Identification of functionally important residues can provide a clear insight into the structural aspects of D2 receptors. In this work, the structure-based approach was applied to identify functionally important residues, while

docking simulations were used to identify the residues involved in the binding pocket. From the earlier docking simulations, the important residues and pharmacophore regions were analyzed by the Ligandscout45 program. Solvated Interaction Energies (SIE) Method. The binding free energy calculations were performed by the SIE method. The SIE method is similar to the linear interaction energy (LIE) methods.46-48 The binding free energies between receptor and inhibitors were calculated for snapshot structures taken from the MD trajectory of the system. From the 4000-ps complex-ligand MD trajectories of the D2 receptor and tetralindiol derivatives inhibitors, 50 snapshots were taken at regular intervals for the binding energy analyses. To provide error estimations for the free energies calculations, we calculated all the data based on the first 25 and last 25 snapshots. The average provided us with a result, and the difference between the two individual values and the average value yielded the estimated error value. The SIE49 function to estimate protein-ligand free energy is written as R ∆Gbind(F, Din, R, γ, C) ) R × [Ec(Din) + ∆Gbind (F, Din) + EVdw + γ∆MSA(F)] + C (1)

where EC and EVdw are the intermolecular Coulomb and van der Waals interaction energies in the bound state, respectively. These values were calculated using the AMBER molecular mechanics force field (FF99) with an optimized dielectric constant. ∆GRbind is the change in the reaction field energy between the bound and free states and is calculated by solving the Poisson equation with the boundary element method program, BRI BEM,50,51 and using a molecular surface generated with a variable-radius solvent probe.52 The ∆MSA term is the change in the molecular surface area upon binding. The following parameters are calibrated by fitting to the absolute binding free energies for a set of 99 proteinsligand complexes: AMBER van der Waals radii linear scaling coefficient (F), the solute interior dielectric constant (Din), the molecular surface area coefficient (γ), the global proportionality coefficient related to the loss of

2372

J. Chem. Inf. Model., Vol. 49, No. 10, 2009

WANG

ET AL.

Figure 3. Ramachandran plot of D2 receptor model 2. Table 4. Reoriented Data of the Important Residues of Model 2 (Dopamine D2 Receptor) from the Docking Simulations compound

hydrogen bondingrelated residues

nonbonding contact-related residues

I II III IV V VI VII

Asp72 Trp60 Thr188 Asp72 Cys76 Null Asp72

Val73 Null Cys76, Met75, Tyr192 Val73, Cys76, Phe187 His175, Leu183, Phe187 Val73, Cys76, Cys167, His175, Leu183, Phe187 Cys76, Leu122, Val130, Trp168, Phe171, His175, Leu183, Phe187

RESULTS AND DISCUSSION

Figure 4. Potential energy plots for compound IIsD2 receptor/ DLPE/water system.

configurational entropy upon binding (R), and a constant (C).46 The optimized values of these parameters are R ) 0.1048, Din ) 2.25, F ) 1.1, γ ) 0.0129 kcal/(mol Å2), and C ) -2.89 kcal/mol. The SIE calculations were carried out with the program sietraj.53

Homology Model Selection, Molecular Docking, and MD Simulation. Initially, two protein models were generatedsmodel 2 being considered the better modelson the basis of a least discrete optimized potential energy (DOPE) score of -23255; this score was found to be very close to that of the other model. The MD simulation (AMBER program) shows that the relaxed model has a DOPE score of -206.526 kcal/mol. Model 2 shows interactions with the seven compounds in the dopamine binding pocket without

DOPAMINE D2 RECEPTOR ANTAGONISTS

J. Chem. Inf. Model., Vol. 49, No. 10, 2009 2373

Figure 5. Analysis of the tetralindiol derivative binding modes. The green and red lines represent hydrogen binding modes. Yellow arcs represent nonbinding contacts.

Figure 6. Calculated binding free energies (error bar) versus experimentally determined binding free energies of the series of seven compounds (each represented by a different colored symbol). The red line represents the regression line.

any docking constraints (Table 3), while model 1 does not show any interactions. Model 2 was also subjected to PROCHECK analysis to assess the stereochemical quality of the model. The Ramachandran plot (Figure 3) shows that 89.5% of residues were found in the most favored and additional allowed regions, 10.0% of the residues were in the generously allowed region, and 0.5% of residues in the

disallowed region. The seven tetralindiol derivative compounds were then docked into model 2; Table 3 lists their best GOLD binding scores. Over the 5-ns MD trajectories of the D2 receptor with solvent molecules and tetralindiol derivatives, the overall structure of both complexes appeared to be equilibrated after 3500 ps. Here, we show the potential energy (Figure 4) of the complex system of compound I. Key Residues of Dopamine D2 Receptors. Our studies of various known compounds have revealed that in the dopamine binding pocket residues can frequently interact with tetralindiol derivatives and that these residues are responsible for the selectivity of tetralindiol derivatives. The analysis results of our simulations are shown in Table 4 and Figure 5. For compound I, Asp72 forms a hydrogen bond with the compound, and Val73 has frequent nonbonding interactions with the compound. For compound II, Trp60 forms a hydrogen bond with the compound. For compound III, Thr188 forms a hydrogen bond with the compound, and the residues Cys76, Met75, and Tyr192 have frequent nonbonding interactions with the compound. For compound IV, Asp72 forms a hydrogen bond with the compound, and the residues Val73, Cys76, and Phe187 have frequent nonbonding interactions with the compound. For compound V, Cys76 forms a hydrogen bond with the compound, and the residues His175, Leu183, and Phe187 have frequent nonbonding interactions with the compound. For compound VI, the residues Val73, Cys76, Cys167, His175, Leu183, and

2374

J. Chem. Inf. Model., Vol. 49, No. 10, 2009

Phe187 have frequent nonbonding interactions with the compound. For compound VII, Asp72 forms a hydrogen bond with the compound, and the residues Cys76, Leu122, Val130, Trp168, Phe171, His175, Leu183, and Phe187 have frequent nonbonding interactions with the compound. The overall results of our simulations suggest that Trp60, Asp72, Cys76, and Thr188 play important roles in forming hydrogen bonds between D2 receptors and tetralindiol derivatives. Moreover, our simulations indicate that Val73, Met75, Cys76, Leu122, Val130, Cys167, Trp168, Phe171, Val173, His175, Leu183, Phe187, and Trp192 have frequent nonbonding interactions with the tetralindiol derivatives. While four residues (Trp60, Asp72, Cys76, and Thr188) were able to form hydrogen bonds between the receptors and the compounds, Asp72 formed the hydrogen bonds most frequently. In addition, Val73, Cys76, Leu183, and Phe187 often formed nonbonding interactions with the compounds. Therefore, our approach theoretically suggests that the five residues Asp72, Val73, Cys76, Leu183, and Phe187 are responsible for the selectivity of the tetralindiol derivatives. Tetralindiol Derivatives Binding Free Energy. The complex structures of D2 receptors with the seven antagonists (compounds I-VII) and the solvent molecules (DLPE lipid bilayer and water molecules) were constructed using the GOLD docking program. The binding free energy of each complex was obtained from the MD simulation and the SIE method, with both processes using the same parameters. All the results are listed in Table 5 in the Supporting Information. In the simulations, the rank order of the predicted binding free energies of the seven compounds is in good agreement with the experimental results (Figure 6), with the correlation coefficient between the predicted and experimental energies being 0.815. This further verifies the predicted D2 receptor structure (model 2) and our docking simulations. CONCLUSION

In this study, we used GOLD, solvent molecules (DLPE lipid bilayer and water molecules), MD simulations, and the SIE method to predict a dopamine D2 receptor structure and a docking model in which a series of tetralindiol derivative antagonists interact with the D2 receptor. In our simulation model, the correlation coefficient between the predicted binding free energies and experimental values of the seven antagonists is 0.815. The Asp72 residue frequently formed hydrogen bonds with the antagonists, and the four residues Val73, Cys76, Leu183, and Phe187 frequently formed nonbonding interactions with the compounds. Therefore, our approach theoretically suggests that the five residues Asp72, Val73, Cys76, Leu183, and Phe187 and the predicted D2 receptor structure can be used as a novel model for an antidopamine D2 receptor to further reconstruct and design new protease antagonists. ACKNOWLEDGMENT

This work was supported by the National Center for Highperformance Computing and National Sun Yat-Sen University, Taiwan. Supporting Information Available: Actual versus predicted data for the tetralindiol derivative antagonists (Table

WANG

ET AL.

5). This material is available free of charge via the Internet at http://pubs.acs.org. REFERENCES AND NOTES (1) Missale, C.; Nash, S. R.; Robinson, S. W.; Jaber, M.; Caron, M. G. Dopamine receptors: from structure to function. Physiol. ReV. 1998, 78, 189–225. (2) Lotharius, J.; Brundin, P. Pathogenesis of parkinson’s disease: dopamine, vesicles and alpha-synuclein. Nature ReV. Neurosci. 2002, 3, 932–942. (3) Kapur, S.; Mamo, D. Half a century of antipsychotics and still a central role for dopamine D2 receptors. Prog. Neuropsychopharm. Bio. Psych. 2003, 27, 1081–1090. (4) Jin, G. Z.; Zhu, A. T.; Fu, Y. Stepholidine: a potential novel antipsychotic drug with dual D1 receptor agonist and D2 receptor antagonist actions. Trends Pharmacol. Sci. 2002, 23, 4–7. (5) Neville, M. J.; Johnstone, E. C.; Walton, R. T. Identification and characterization of ANKK1: a novel kinase gene closely linked to DRD2 on chromosome band 11q23.1. Hum. Mutat. 2004, 23, 540– 545. (6) Payne, S. L.; Johansson, A. M.; Strange, P. G. Mechanisms of ligand binding and efficacy at the human D2(short) dopamine receptor. J. Neurochem. 2002, 82, 1106–1117. (7) Javitch, J. A.; Fu, D.; Chen, J.; Karlin, A. Mapping the binding-site crevice of the dopamine D2 receptor by the substitutedcysteine accessibility method. Neuron 1995, 14, 825–831. (8) Grundt, P.; Husband, S. L. J.; Luedtke, R. R.; Taylor, M.; Newman, A. H. Analogues of the dopamine D2 receptor antagonist L741,626: binding, function, and SAR. Bioorg. Med. Chem. Lett. 2007, 17, 745– 749. (9) Imhof, R.; Godel, T.; Burkard, W. P.; Keller, H. H.; Da, P. M.; Scherschilicht, R. New insight into structural and stereochemical requirements for selective, high affinity ligands at the dopamine D2 receptor. Neurochem. Int. 1992, 20, 75S–80S. (10) Naijue, Z.; Li, L.; Stevens, C. L. K. A CoMFA study of Dopamine D2 Receptor Agonists and X-Ray Crystal Structure of Quinelorane Dihydrochloride Dihydrate, R(-)-Apomorphine Hydrochloride and R(-)-N-n-Propylnorapomorphine Hydrochloride. Struct. Chem. 2004, 6, 553–565. (11) Chidester, C. G.; Lin, C. H.; Lahti, R. A.; Haadsma-Svensson, S. R.; Smith, M. W. Comparison of 5-HT1A and dopamine D2 pharmacophores. X-ray structures and affinities of conformationally constrained ligands. J. Med. Chem. 1993, 36, 1301–1315. (12) Johansson, A. M.; Karlen, A.; Grol, C. J.; Sundell, S.; Kenne, L.; Hacksell, U. J. Dopaminergic 2-aminotetralins: affinities for dopamine D2-receptors, molecular structures, and conformational preferences. Mol. Pharmacol. 1986, 30, 258–269. (13) Gro¨zinger, M.; Dragicevic, A.; Hiemke, C.; Shams, M.; Mu¨ller, M. J.; Ha¨rtter, S. Melperone is an inhibitor of the CYP2D6 catalyzed O-demethylation of venlafaxine. Pharmacopsychiatry 2003, 36, 3–6. (14) Szarfman, A.; Tonning, J. M.; Levine, J. G.; Doraiswamy, P. M. Atypical antipsychotics and pituitary tumors: a pharmacovigilance study. Pharmacotherapy 2006, 26, 748–58. (15) Seeman, P.; Watanabe, M.; Grigoriadis, D.; Tedesco, J. L.; George, S. R.; Svensson, U.; Nisson, J. L. G.; Neumeyer, J. L. Dopamine D2 receptor binding sites for agonists. Mol. Pharmacol. 1985, 28, 391– 399. (16) Mewshaw, R. E.; Verwijs, A.; Shi, X.; McGaughey, G. B.; Nelson, J. A.; Mazandarani, H.; Brennan, J. A.; Marquis, K. L.; Coupet, J.; Andree, T. H. New generation dopaminergic agents. 5. heterocyclic bioisosteres that exploit the 3-OH-N1-phenylpiperazine dopaminergic template. Bioorg. Med. Chem. Lett. 1998, 8, 2675–2680. (17) Martin, Y. C.; Bures, M. G.; Dabaher, E. A.; Delazzer, J.; Lico, I.; Pavlik, P. A. A fast new approach to pharmacophore mapping and its application to dopaminergic and benzodiazepine agonists. J. Comput.Aided Mol. Des. 1993, 7, 83–102. (18) Newman, A. H.; Grundt, P.; Cyriac, G.; Deschamps, J. R.; Taylor, M.; Kumar, R.; Ho, D.; Luedtke, R. R. N-(4-(4-(2,3-Dichloro-or2methoxyphenyl)piperazin-1-yl)butyl)heterobiarylcarboxamides with functionalized linking chains as high affinity and enantioselective D3 receptor antagonists. J. Med. Chem. 2009, 52, 2559–2570. (19) Fu, W.; Shen, J.; Luo, X.; Zhu, W.; Cheng, J.; Yu, K.; Briggs, J. M.; Jin, G.; Chen, K.; Jiang, H. Dopamine D1 Receptor Agonist and D2 Receptor Antagonist Effects of the Natural Product (2)sStepholidine: Molecular Modeling and Dynamics Simulations. Biophys. J. 2007, 93, 1431–1441. (20) Nervall, M.; Hanspers, P.; Carlsson, J.; Boukharta, L.; Åqvist, J. Predicting binding modes from free energy calculations. J. Med. Chem. 2008, 51, 2657–2667. (21) Bertolino, A.; Fazio, L.; Giorgio, A.; Blasi, G.; Romano, R.; Taurisano, P.; Caforio, G.; Sinibaldi, L.; Ursini, G.; Popolizio, T.; Tirotta, E.;

DOPAMINE D2 RECEPTOR ANTAGONISTS

(22)

(23)

(24)

(25)

(26)

(27)

(28) (29) (30) (31)

(32) (33) (34) (35) (36)

Papp, A.; Dallapiccola, B.; Borrelli, E.; Sadee, W. Genetically Determined Interaction between the Dopamine Transporter and the D2 Receptor on Prefronto-Striatal Activity and Volume in Humans. J. Neurosci. 2009, 29, 1224–1234. Altschul, S. F.; Madden, T. L.; Schaffer, A. A.; Zhang, J. H.; Zhang, Z.; Miller, W.; Lipman, J. T. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 3389–3402. Hanson, M. A.; Cherezov, V.; Griffith, M. T.; Roth, C. B.; Jaakola, V. P.; Chien, E. Y.; Velasquez, J.; Kuhn, P.; Stevens, R. C. A specific cholesterol binding site is established by the 2.8 A structure of the human beta2-adrenergic receptor. Structure 2008, 16, 897–905. Rosenbaum, D. M.; Cherezov, V.; Hanson, M. A.; Rasmussen, S. G.; Thian, F. S.; Kobilka, T. S.; Choi, H. J.; Yao, X. Y.; Weis, W. I.; Stevens, R. C.; Kobilka, B. K. GPCR engineering yields highresolution structural insights into beta2-adrenergic receptor function. Science 2007, 318, 1266–1273. Jaakola, V. P.; Griffith, M. T.; Hanson, M. A.; Cherezov, V.; Chien, E. Y.; Lane, J. R.; Ijzerman, A. P.; Stevens, R. C. The 2.6 a crystal structure of a human a2a adenosine receptor bound to zm241385. Science 2008, 322, 1211–1217. Rasmussen, S. G.; Choi, H. T.; Rosenbaum, D. M.; Kobilka, T. S.; Thian, F. S.; Edwards, P. C.; Burghammer, M.; Ratnala, V. R.; anishvili, R.; Fischetti, R. F.; Schertler, G. F.; Weis, W. I.; Kobilka, B. K. Crystal structure of the human beta2 adrenergic G-proteincoupled receptor. Nature 2007, 450, 383–387. Thompson, J. D.; Gibson, T. J.; Plewniak, F.; Jeanmougin, F.; Higgins, D. G. The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res. 1997, 25, 4876–4782. Marti-Renom, M. A.; Stuart, A.; Fiser, A.; Sa´nchez, R.; Melo, F.; Sali, A. Comparative protein structure modeling of genes and genomes. Annu. ReV. Biophys. Biomol. Struct. 2000, 29, 291–325. Sali, A.; Blundell, T. L. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 1993, 234, 779–815. Fiser, A.; Do, R. K.; Sali, A. Modeling of loops in protein structures. Protein Sci. 2000, 9, 1753–1773. Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. The Amber biomolecular simulation programs. J. Comput. Chem. 2005, 26, 1668– 1688. Laskowski, R. A.; MacArthur, M. W.; Moss, D. S.; Thornton, J. M. PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Crystallogr. 1993, 26, 283–291. SYBYL, Version 8.1; Tripos International: St. Louis, MO, 2008. Jones, G.; Willett, P.; Glen, R. C.; Leach, A. R.; Taylor, R. Development and Validation of a Genetic Algorithm for Flexible Docking. J. Mol. Biol. 1997, 267, 727–748. Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and testing of a general Amber force field. J. Comput. Chem. 2004, 25, 1157–1174. Bren, U.; Hodoscek, M.; Koller, J. Development and validation of empirical force field parameters for netropsin. J. Chem. Inf. Model. 2005, 45, 1546–1552.

J. Chem. Inf. Model., Vol. 49, No. 10, 2009 2375 (37) Vailikhita, V.; Treesuwana, W.; Hannongbua, S. A combined MDsONIOM2 approach for 1H NMR chemical shift calculations including a polar solvent. J. Mol. Struct.: THEOCHEM. 2007, 806, 99–104. (38) Lu, Q.; Tan, Y. H.; Luo, R. Molecular dynamics simulations of p53 DNA-binding domain. J. Phys. Chem. B 2007, 111, 11538–11545. (39) Li, W.; Ode, H.; Hoshino, T.; Liu, H.; Tang, Y.; Jiang, H. Reduced catalytic activity of P450 2A6 mutants with Coumarin: a computational investigation. J. Chem. Theory Comput. 2009, 5, 411–1420. (40) Olson, A. J.; Hu, Y. H.; Keinan, E. Chemical mimicry of viral capsid self-assembly. Proc. Natl. Acad. Sci. 2007, 104, 20731–20736. (41) Gaussian 03, Version C.02; Gaussian, Inc.: Wallingford, CT, 2004. (42) Wang, J. M.; Wang, W.; Kollman, P. A. Antechamber: An accessory software package for molecular mechanical calculations. Abstr. Pap. Am. Chem. Soc. 2001, 222, U403–U403. (43) Ryckaert, J. P.; Ciccotti, G. Andersen’s canonical-ensemble molecular dynamics for molecules with constraints. Mol. Phys. 1986, 58, 1125– 1136. (44) Ryckaert, J. P.; Ciccotti, G.; Berendsen, H. J. C. integration of the cartesian equations of motion of a system with constrains: molecular dynamics of n-alkanes. J. Comput. Phys. 1977, 23, 327–341. (45) Wolber, G.; Dornhofer, A. A.; Langer, T. Efficient overlay of small organic molecules using 3D pharmacophores. J. Comput.-Aided Mol. Des. 2007, 20, 773–788. (46) Perdih, A.; Bren, U.; Solmajer, T. Binding free energy calculations of N-sulphonyl-glutamic acid inhibitors of MurD ligase. J. Mol. Model. 2009, 15, 983–996. (47) Bren, U.; Martı´nek, V.; Floria´n, J. Free energy simulations of uncatalyzed DNA replication fidelity: structure and stability of T · G and dTTP · G terminal DNA mismatches flanked by a single dangling nucleotide. J. Phys. Chem. B 2006, 110, 10557–10566. (48) Åqvist, J.; Medina, C.; Samuelsson, J. E. A new method for predicting binding affinity in computer-aided drug design. Protein Eng. 1994, 7, 385–391. (49) Naı¨m, M.; Bhat, S.; Rankin, K. N.; Dennis, S.; Chowdhury, S. F.; Siddiqi, I.; Drabik, P.; Sulea, T.; Bayly, C.; Jakalian, A.; Purisima, E. O. Solvated interaction energy (SIE) for scoring protein-ligand binding affinities. 1. Exploring the parameter space. J. Chem. Inf. Model. 2007, 47, 122–133. (50) Purisima, E. O. Fast Summation Boundary Element Method for Calculating Solvation Free Energies of Macromolecules. J. Comput. Chem. 1998, 19, 1494–1504. (51) Purisima, E. O.; Nilar, S. H. A. simple yet accurate boundary element method for continuum dielectric calculations. J. Comput. Chem. 1995, 16, 681–689. (52) Bhat, S.; Purisima, E. O. Molecular surface generation using a variableradius solvent probe. Proteins: Struct., Funct., Bioinf. 2006, 62, 244– 261. (53) Cui, Q.; Sulea, T.; Schrag, J. D.; Munger, C.; Hung, M.-N.; Naı¨m, M.; Cygler, M.; Purisima, E. O. Molecular Dynamics and Solvated Interaction Energy Studies of Protein-Protein Interactions: the MP1p14 Scaffolding Complex. J. Mol. Biol. 2008, 379, 787–802.

CI9002238